You are on page 1of 7

Chemical Engineering and Processing 50 (2011) 338344

Contents lists available at ScienceDirect

Chemical Engineering and Processing:


Process Intensication
journal homepage: www.elsevier.com/locate/cep

Electro-coalescence of an aqueous droplet at an oilwater interface


M. Mousavichoubeh a,b , M. Ghadiri a, , M. Shariaty-Niassar b
a
b

Institute of Particle Science and Engineering, University of Leeds, Leeds LS2 9JT, UK
School of Chemical Engineering, College of Engineering, University of Tehran, 11365-4563 Tehran, Iran

a r t i c l e

i n f o

Article history:
Received 1 April 2010
Accepted 21 September 2010
Available online 29 September 2010
Keywords:
Separation
Electric eld
Electro-coalescence
Oilwater interface
Secondary droplet

a b s t r a c t
The coalescence of an aqueous droplet at an oilwater interface under an electric eld has been investigated, with a view to quantify conditions that give rise to secondary droplet formation. Two patterns
of drop-interface coalescence may occur: complete coalescence and partial coalescence. The former is
obviously the desirable pattern for industrial coalescers. However in practice, the process of coalescence
could actually produce smaller droplets, which become more difcult to remove, and hence undesirable.
This is caused by either necking, due to extensive elongation of the droplet, or reaction to a fast and energetic coalescence and is referred to as partial coalescence. The volume of the droplets formed in this way
has been analyzed as a function of the initial droplet size, electric eld strength and the distance between
the droplet and the interface. The expansion speed of the neck connecting the droplet and interface at
the beginning of the pumping process has also been quantied. These results are useful in optimizing the
electro-coalescence process.
2010 Elsevier B.V. All rights reserved.

1. Introduction
In the chemical, processing and manufacturing industries,
immiscible liquids are often mixed such that one phase is fully dispersed in another, e.g. in extraction and leaching. The aim is to get a
large interfacial area for the enhancement of mass transfer between
the two immiscible liquids [14]. In crude oil extraction from oil
wells [5], an aqueous saline phase is often well-dispersed in the
crude oil. However these emulsions or dispersions have to be separated into their constituent phases before the next operating steps
or as required by process requirements, environmental regulations
and customer specications as in the case of crude oil industry [6].
There are several techniques for enhancing the separation of waterin-oil emulsions, such as the addition of chemical demulsier [7],
pH adjustment and ltration [8], gravity or centrifugal settling [9],
heat treatment and electrostatic demulsication [10,11]. In terms
of energy efciency, electrostatic demulsication is considered to
be the best among the above methods [10]. The state of the art
has been reviewed by Eow and Ghadiri [6]. Here the rate of coalescence can signicantly be enhanced by the application of an electric
eld. The coalescence occurs in three stages [9,12,13]. In the rst
stage, the drops approach each other or the interface and are separated by a lm of the continuous phase. The second stage involves
the thinning of this lm. When the lm reaches a critical thickness
any disturbance or instability causes it to rupture, following which

Corresponding author. Tel.: +44 113 343 2406; fax: +44 113 343 2384.
E-mail address: m.ghadiri@leeds.ac.uk (M. Ghadiri).
0255-2701/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2010.09.017

coalescence occurs [1416]. Film thinning is often the overall controlling step in the absence of an electric eld. In order to increase
the separation rate, the lm-thinning process needs to be faster and
this can be done by the application of an electric eld. High electric
elds have been used to separate water-in-oil dispersions in crude
oil and extraction industries [6]. To apply this method, the continuous phase needs to be much more electrically insulating compared
with the dispersed phase, in order to set up an electric eld [17].
The current understanding of the electrocoalscence phenomenon
has been reviewed by Eow et al. [18].
2. The effect of applied electric eld on drop interface
coalescence
In the absence of an electric eld the coalescence of a drop
at a liquidliquid interface sometimes produces smaller drops as
observed by Charles and Mason [19,20]. During the rupture of
the lm between the main drop and the interface, the excess
internal pressure from the curved interface produces a cylindrical liquid column. The radius of this column decreases rapidly until
its circumference becomes smaller than its height. As a result of
a Rayleigh wave disturbance, a secondary droplet is often formed
[20]. The mean rest-time of a drop at an interface can be significantly reduced by applying an electric eld because the rate of
lm thinning is increased [1922]. It has been reported that in the
presence of an external electric eld, the generation of the secondary droplet does not occur [19,20,22]. Aryafar and Kavehpour
[23] have recently proposed that the partial coalescence in the
absence of an electric eld may be described by Ohnesorge number,

M. Mousavichoubeh et al. / Chemical Engineering and Processing 50 (2011) 338344

339

GlassmanHigh Voltage Inc.). The bottom electrode was grounded.


A high-speed digital video camera (Photron FASTCAM SA5 32GB),
equipped with a micro lens (NAVITARA 12X Zoom Lens) was used
to observe the phenomena taking place during the drop-interface
coalescence. Using this camera a framing speed of up to 62,000 fps
was used to record the coalescence process. The video camera
was focused on the centre of the liquidliquid interface. A halogen
lamp (DedolightDLHM4-300) with four exible ber optic heads
was used for lighting. The intensity of the lighting could be accurately adjusted to facilitate focusing. De-ionized water was used
to produce the droplets, while the organic phase was sunower
oil (obtained from Morrisons, UK Ltd.). The properties of the liquids used in this research are given in Table 1. The conductivity of
the liquids was measured using a conductivity meter, Model 4310
from Jenway Products Inc. Liquid viscosity was measured using a
shear rate and yield stress rheometer, Bohlin CVO Rheometer from
Malvern Instruments Inc. The viscosity does not change with the
shear rate, as the continuous uid is incompressible Newtonian
liquid. The density was measured using a volumetric ask. The
dielectric constant of the liquids was obtained from the literatures
[26] and interfacial tensions were measured according to a technique based on the pendant drop method, using a contact-angle
measuring instrument, EasyDrop from Kruss GmbH.
The densities of sunower oil and de-ionized water are quite
close and this facilitates experimentation, because it allows water
droplets falling slowly towards the liquidliquid interface, giving
ample time for managing the recording process. Water droplets of
different sizes were produced using hypodermic needles by pressing the syringe piston so that it changed the volume of the uid
inside the syringe and by this method it was possible to have a
droplet diameter range between 576 1 m and 1196 4 m. The
glassware and needles were cleaned by several washing using acetone and distilled water to prevent any pollution that may affect
the coalescence phenomenon. The droplets were released gently
from the needle to avoid oscillation of the droplets and interface.
They were then introduced at the middle of the planar interface,
which was far away from the edge, where the interface was curved
and the electric eld was not uniform. The diameter of the droplets
was measured with an accuracy of 5 m by the use of ImagePro and PFV (Photron Fastcam Viewer) software. The diameter of
the needle was measured by a microscope and was used for the
in situ calibration of the droplet size. The experiments were done
at 21 2 C.

Fig. 1. The Prespex test-cell [25].

Oh = /(R 12 )0.5 , where  is viscosity, R is the radius of the drop,


 is density,  is interfacial surface tension, and the indices 1 and 2
indicate the medium and drop properties, respectively. For Oh > 1
full coalescence occurs and for Oh < 1only partial coalescence takes
place. They have also analysed the effect of electric elds on coalescence [24] where they report that partial coalescence is suppressed
under electric elds in the high Ohnesorge number domain.
3. Objectives of the present work
For the separation of water-in-oil emulsions, the smaller the dispersed phase droplet size is, the more difcult would the separation
be. An important problem in this separation process is the formation of the secondary droplet by partial coalescence. This leads to
a lowering of the separation efciency as the much tinier droplets
are more difcult to separate. Therefore basically it is better to prevent these secondary droplets from forming in the rst place. In this
work our observations on the formation of the secondary droplets
during the electro-coalescence of the primary droplets are reported
and the parameters that affect the process are quantied.
4. Experimental set-up and procedure
The experimental cell used in this work is shown in Fig. 1.The cell
was made of Perspex to facilitate visualization of the phenomenon.
The electrodes were polished brass plates. The plates had dimensions of 90 mm 25 mm.The high voltage electrode was attached
to the moveable upper part of the cell. The distance between the
two electrodes could therefore be varied by moving the upper part
up or down although it was set at 53 mm in this work.
The Perspex block had a thickness of about 6 mm.There is a small
hole through the middle point of the moveable upper part of the
cell and the brass plate for a hypodermic needle to go through it.
The needle attached to a syringe (Hamilton micro-liter syringe)
was used to produce small aqueous droplets in the cell. The high
voltage electrode was connected to a positive polarity high voltage
direct current source (Model: PS/EH60R01.5-22, manufactured by

5. Results and discussion


5.1. The mechanism of drop-interface coalescence under electric
eld
The coalescence events of water droplets at an oil/water interface were recorded at 20,000 fps and are shown in the sequence
of images in Figs. 2 and 3. In these gures, the length scale is
the same for all images, facilitating the comparison between the
primary droplet and the secondary droplets. The deformation of
the droplet and interface before coalescence in the absence of an
electric eld has been reported previously by Charles and Mason
[20], but this is notable only when the drop is resting on the
interface. Applying a sufciently strong electric eld results in the

Table 1
The properties of the liquid used in the experiment.
Liquids

Conductivity (S m1 ) (5%)

Viscosity (mPa s) (5%)

Surface tensiona (mN m1 ) (5%)

Density (kg m3 ) (5%)

Dielectric constant

De-ionized water
Sunower oil

1.1 10
3.6 106

1.00
46.5

73
33

998
920

80
4.9

The measured surface tension is with respect to air at 1 atm and 20 C.

340

M. Mousavichoubeh et al. / Chemical Engineering and Processing 50 (2011) 338344

Fig. 2. Sequence of a complete coalescence for a droplet of 984 2 m diameter under electric eld strength of 56 V/mm.

local deformation of interface and falling droplet (see Fig. 4(b)). An


uncharged droplet subject to an electric eld is polarized and when
the deformed droplet approaching the deformed interface the electric eld strength increases exponentially at small separations to
the extent that electroclamping phenomenon becomes operative
[27] giving rise to the neck formation as shown in Fig. 4(b). The
high speed of video recording has made this observation very
clear. When the droplet is sufciently close to the interface (a
few micrometers apart) in a time period less than 16 s the high
strength electric eld in the gap between the droplet and interface
causes clamping between them (Fig. 4(b)), resulting in the formation of a narrow channel. The droplet will now be acquiring the
same charge as the adjacent electrode and will be experiencing a
repelling Columbic force. For a solid electrode plate, the droplet
will be repelled from the electrode. However for an aqueous liquid interface, the surface tension will be pushing the liquid in the
droplet into the continuous phase via the channel formed, and
rapidly enlarging the neck. It is likely that the current constriction
of the electrical clamping process ruptures the thin lm (oil phase)
between the droplet and the interface. A hole is formed in this way
at the interface by which the process of coalescence is initiated.
The hole expands very rapidly and the liquid is pumped rapidly
into its bulk phase. Two patterns of coalescence are observed here:
complete coalescence and partial coalescence, as shown in
Figs. 2 and 3, respectively.
There are two rate processes operating: pumping of droplet into
its bulk phase (due to surface tension) and the necking process.
Whether a secondary droplet is formed depends on the process
which is dominating. The predominance of each of these processes,
i.e. necking and pumping depends on some parameters and will be
discussed later.
5.2. Complete and partial coalescence patterns
It can be seen from Fig. 2 that with the start of droplet pumping
into its bulk phase, no necking occurs leading to a complete coalescence pattern. In this coalescence pattern in some stages (Fig. 2(l)
and (m)), the peak of the droplet becomes sharp but nally as a
result of surface tension the process leads to a at surface without any detached bodies. This means in such a condition the rate

of pumping is faster than that of the necking process. Although


the droplet experiences a repelling Columbic force by the adjacent electrode, this force is not sufciently strong (in the case of
complete coalescence) to help the occurrence of necking process.
Fig. 3 shows the pattern of coalescence for droplets subjected to
a higher electric eld strength than that of Fig. 2, hence a necking
process is observed. In Fig. 3(b), as a result of a faster development
of the necking process in comparison with Fig. 3(a), the volume
of the secondary droplet is bigger. In case (b) the repelling force
between the droplet and the interface (after charging) is stronger
than that for case (a) because of larger electric eld strength and
the necking process develops faster detaching a bigger secondary
droplet. In both cases (a) and (b) the pumping process is in progress
to drain the content of the primary droplet into its bulk phase
but the necking process and its rate determine the volume of
the secondary droplet. Thus under a given condition leading to
a partial coalescence, the necking process and the pumping process are in competition with each other. Another noteworthy point
is the difference between the tail length in Fig. 3(a) and (b). In
Fig. 3(b) the droplet is subjected to higher electric eld strength
and the detached body has a longer tail than the other case. The
secondary droplet is moving upwards, stretching the neck even
further before it is broken into several much ne droplets. In conclusion these two strong electric elds are detrimental for efcient
electro-coalescers.
5.3. Factors affecting the volume of the secondary droplet
High speed video observations indicate that three parameters
are inuential in determining the coalescence pattern: droplet size
(d), electric eld strength (E) and the height of falling droplet from
interface (). In this work the dispersed phase is de-ionized water
and does not contain any surfactants which could affect the interfacial tension.
5.3.1. The effect of primary droplet size
The effect of droplet size in the range 576 11196 4 m on
the volume of the secondary droplet has been investigated for
constant electric eld strength and . The results are shown in
Fig. 5(a)(d).

M. Mousavichoubeh et al. / Chemical Engineering and Processing 50 (2011) 338344

341

Fig. 3. Partial coalescence of a droplet of 1196 4 m diameter under two electric eld strengths: (a) 124 V/mm and (b) 181 V/mm.

As observed from these gures, by increasing the primary


droplet size (d) the volume of detached body increases, except for
 = 0 (not shown here) where the coalescence is complete. It means
when the droplet is sitting on the interface there will not be any
detached body and all coalescence processes lead to a complete
coalescence pattern. The formation of bigger detached bodies as
the primary droplet size is increased can be explained by considering the effect of interfacial tension. According to Young-Laplace,
p = 2/R, where p,  and R are internal pressure, interfacial
tension and the droplet radius, the bigger droplets have lower internal pressure, as compared with the smaller ones [28]. They will
therefore be more deformable and form a long neck more readily.
Another factor is that the polarization of larger droplets can be more
effective in producing a neck. Thus less rigidity and larger attractive
force by the far electrode on a bigger droplet result in detaching a
bigger volume from primary droplets.

Fig. 4. Deformation and start of coalescence of a 1196 4 m droplet with an interface under the electric eld strength of 181 V/mm, showing the formation of a
narrow channel.

5.3.2. The effect of electric eld strength


The results of Fig. 5(a)(d) can be re-plotted in terms of the electric eld strength to show its trend, one of which is shown in Fig. 6.
It can be seen that when the falling droplet is subjected to higher

M. Mousavichoubeh et al. / Chemical Engineering and Processing 50 (2011) 338344

(a)

56 V/mm

90 V/mm

124 V/mm

158 V/mm

181 V/mm

Detached body vol103 (mm3)

4
3.5
3
2.5
2
1.5
1
0.5
0
500

600

700

800

900

1000

1100

1200

(b)
Detached body vol103 (mm3)

342

56 V/mm

90 V/mm

90 V/mm

124 V/mm

158 V/mm

181 V/mm

35
30
25
20
15
10
5
0
500

600

700

800

181 V/mm

8
6
4
2
0
500

600

700

800

900

1000

1100

1200

d (m)

(d)
Detached body vol103 (mm3)

Detached body vol103 (mm3)

56 V/mm

158 V/mm

10

d (m)

(c)

124 V/mm

12

900

1000

1100

1200

d (m)

56 V/mm

90 V/mm

124 V/mm

158 V/mm

181 V/mm

70
60
50
40
30
20
10
0
500

600

700

800

900

1000

1100

1200

d (m)

Fig. 5. (a)(d) The effect of droplet size on detached body volume under various electric eld strengths and separation distances: (a)  = 53 7 m, (b)  = 101 10 m, (c)
 = 150 11 m, and (d)  = 200 14 m

Detached body vol103 (mm3)

5761m

8291m

9842m

11002m

11964m

12

Table 2
The deformation of the droplet diameter of 984 2 m under different electric eld
strength quantitatively.

10
8

Electric eld strength (V/mm)

90

124

158

181

D = dmajor /dminor

1.058

1.075

1.154

1.158

6
4
2
0
50

70

90

110

130

150

170

190

Electric Field Strength (V/mm)


Fig. 6. The effect of electric eld strengths on detached body volume for various
drop size and  = 101 10 m.

electric eld strengths the volume of the detached body increases.


As mentioned, the volume of detached body is being controlled
by the mutual interaction between the necking and pumping process. Eow et al. [25] have shown that by increasing the electric eld
strength the deformation of a given droplet size increases.

As shown in Fig. 7 and Table 2, by increasing the strength of


electric eld from 90 V/mm to 181 V/mm the deformation of the
droplet with the diameter of 984 2 m increases and that the side
of droplet close to the interface gets more elongated locally under
higher electric elds. This has previously been shown by Eow et al.
[25]. The deformations in Table 2 are the ratio of the major to minor
diameters of the drop.
As the narrow channel forms between droplet and interface
the pumping process competes with the necking, with the latter
brought about by the change in droplet polarity, due to contact, and
consequent repelling from the adjacent electrode. So the pumping
process is not able to overcome the necking process. In this state and
for a given droplet size and falling height, the volume of detached
body in a higher electric eld will be bigger than that for lower elec-

Fig. 7. Different deformations degree for the same droplet diameter of 984 2 m under different electric eld strength.

M. Mousavichoubeh et al. / Chemical Engineering and Processing 50 (2011) 338344

Detached body vol103(mm3)

5761m

8291m

343

9842m

11002m

11964m

70
60
50
40
30
20
10
0
0

25

50

75

100

125

150

175

200

225

(m)
Fig. 10. The effect of height of drop from interface on detached body volume for
various drop sizes under E = 181 V/mm.

6. Conclusions
Fig. 8. Characteristic dimensions to monitoring the speeds.

tric elds. The temporal changes in some characteristic dimensions


during coalescence are shown in Figs. 8 and 9.
In Fig. 9, channel expansion radial speed shows the radial
change (drchannel (t)/dt) of the narrow channel connecting the
droplet and interface with time, and the terms equator reduction
(drequator (t)/dt) and peak falling (dhpeak (t)/dt) show the radial
reduction and falling of peak point of droplet respectively. To
understand the pumping behavior of the droplet these three speeds
should be considered at the same time. As it can be seen in Fig. 9, at
the beginning of the coalescence by formation of the narrow channel between droplet and interface, the speed of channel expansion
is very fast (in this case about 245 mm/s) and it decreases continuously. Moreover, the speeds of equator reduction and peak falling
are initially zero.
5.3.3. The effect of height of droplet from interface ()
The data of Fig. 5(a)(d) can be expressed in terms of  to see its
effect on secondary droplet volume clearly. This is done for one
electric eld strength in Fig. 10. Increasing the height of falling
droplet from interface for a given droplet size and under a constant
electric eld strength, the volume of detached body increases.
The droplets located further from the interface, i.e. the larger
value of , experience stronger attractive Columbic forces (F 1/r2
where r is the distance between charge sources) exerted by the high
potential electrode. So less rigidity and larger Columbic force helps
necking process signicantly.
Channel Expansion

Equator Reducon

Peak Falling

245
220

Radial Speed (mm/s)

195
170
145
120
95
70
45
20
-5
-30
-55

10

11

12

t (ms)

Fig. 9. Speed of the channel expansion, equator reduction and peak falling for a drop
with diameter 984 2 m under E = 181 V/mm.

Drop-interface coalescence under an electric eld could lead to


the formation of a detached body. This is of course highly undesirable, and should be avoided by optimizing the process. The
parameters that affect the formation of detached body have been
analyzed. These are the droplet size, electric eld strength and
distance of droplets from the electrodes. The results show that
by increasing any of these parameters the volume of detached
body increases. This information is useful in optimizing the electrocoalescence process. For example, by the use of appropriate
pulsatile electric eld it could be possible to enhance the electrocoalescence and at the same time suppress the formation of
secondary droplets.
Acknowledgements
The visit of M. Mousavichoubeh to the University of Leeds was
facilitated by nancial support from the National Iranian Oil Company. Due to the short-term nature of the visit, the work could
not be done without generous help from a large number of members of Institute of Particle Science and Engineering. Thanks are
especially due to Dr. Ali Hassanpour and Messrs Graham Calvert,
Massih Pasha, Colin Hare, Robert Harris, Javad Khangostar, Tarsem
Hunjan, Nejat Rahmanianand Miss Liming Zhang. The authors are
also grateful for the support of the project co-supervisors, Professors H. Bahmanyar and M. A. Moosavian, College of Engineering,
University of Tehran.
References
[1] R.H. Perry, D.W. Green, J.O. Maloney, Perrys Chemical Engineers Handbook,
7th ed., McGraw-Hill, New York, 1997.
[2] T.M. Dreher, J. Glass, A.J. OConnor, G.W. Stevens, Effect of rheology on coalescence rates and emulsion stability, AIChE Journal 45 (6) (1999) 11821190.
[3] S. Hartland, B. Yang, S.A.K. Jeelani, Dimple formation in the thin lm beneath a
drop or bubble approaching a plane surface, Chemical Engineering Science 49
(9) (1994) 13131322.
[4] J.C. Godfrey, M.J. Slater, LiquidLiquid Extraction Equipment, Wiley, London,
1994.
[5] D.G. Thompson, A.S. Taylor, D.E. Graham, Emulsication and demulsication
related to crude oil production, Colloids and Surfaces 15 (1985) 175189.
[6] J.S. Eow, M. Ghadiri, Electrostatic enhancement of coalescence of water droplets
in oil: a review of the technology, Chemical Engineering Journal 85 (23) (2002)
357368.
[7] R.A. Mohammed, A.I. Baile, P.F. Luckham, S.E. Taylor, Dewatering of crude oil
emulsions: 3. Emulsion resolution by chemical means, Colloids and Surfaces A:
Physicochemical and Engineering Aspects 83 (3) (1994) 261271.
[8] K.J. Lissant, Demulsication: Industrial Application, Surfactant Science Series,
vol. 13, Marcel Dekker, New York, 1983.
[9] D. Sun, S.C. Jong, X.D. Duan, D. Zhou, Demulsication of water-in-oil emulsion
by wetting coalescence materials in stirred- and packed-columns, Colloids and
Surfaces A 150 (1999) 6975.
[10] M. Goto, J. Irie, K. Kondo, F. Nakashio, Electrical demulsication of w/o emulsion
by continuous tubular coalescer, Journal of Chemical Engineering of Japan 22
(4) (1989) 401406.

344

M. Mousavichoubeh et al. / Chemical Engineering and Processing 50 (2011) 338344

[11] R.A. Mohammed, A.I. Baile, P.F. Luckham, S.E. Taylor, Dewatering of crude oil
emulsions 1. Rheological behaviour of the crude oilwater interface, Colloids
and Surfaces A: Physicochemical and Engineering Aspects 80 (23) (1993)
223235.
[12] D. Sun, X. Duan, W. Li, D. Zhou, Demulsication of water-in-oil emulsion by
using porous glass membrane, Journal of Membrane Science 146 (1998) 6572.
[13] C.T. Chen, J.R. Maa, Y.M. Yang, C.H. Chang, Effects of electrolytes and polarity of
organic liquids on the coalescence of droplets at aqueousorganic interfaces,
Surface Science 406 (1998) 167177.
[14] E.E. Isaacs, R.S. Chow, Practical aspects of emulsion stability, Advances in Chemistry Series 231 (1992) 251277.
[15] S.E. Friberg, S. Jones, Emulsions, in: Kirk-Othmer Encyclopedia of Chemical Technology, vol. 9, 4th ed., John Wiley & Sons, Inc., 1996, pp. 393
413.
[16] J. Holto, G. Berg, L.E. Lundgaard, Electrocoalescence of drops in a water-in-oil
emulsion, in: Electrical Insulation and Dielectric Phenomena, 2009. CEIDP 09,
IEEE Conference on, October 1821, 2009,Virginia Beach, VA, IEEE, 2010, pp.
196199.
[17] L.E. Lundgaard, G. Berg, S. Ingebrigtsen, P. Atten, Electrocoalescence for
oilwater separation: fundamental aspects, in: J. Sjblom (Ed.), Emulsion and
Emulsion Stability, 2nd ed., Taylor & Francis Group, LLC, Boca Raton, 2006, pp.
549592.
[18] J.S. Eow, M. Ghadiri, A.O. Sharif, T.J. Williams, Electrostatic enhancement of
coalescence of water droplets in oil: a review of the current understanding,
Chemical Engineering Journal 84 (3) (2001) 173192.

[19] G.E. Charles, S.G. Mason, The coalescence of liquid drops with at liquid/liquid
interfaces, Journal of Colloid Science 15 (3) (1960) 236267.
[20] G.E. Charles, S.G. Mason, The mechanism of partial coalescence of liquid drops
at liquid/liquid interfaces, Journal of Colloid Science 15 (2) (1960) 105122.
[21] G.A.H. Elton, R.G. Picknett, The coalescence of aqueous droplets with an oilwater
interface, Proceedings of International Congress on Surface Activity B (1957)
288294.
[22] R.S. Allan, S.G. Mason, Effects of electric elds on coalescence in liquidliquid
system, Transactions of the Faraday Society 57 (1961) 2027.
[23] H. Aryafar, H.P. Kavehpour, Drop coalescence through planar surfaces, Physics
of Fluids 18 (7) (2006) 07210572106.
[24] H. Aryafar, H.P. Kavehpour, Electrocoalescence:, Effects of DC electric elds
on coalescence of drops at planar interfaces, Langmuir 25 (21) (2009)
1246012465.
[25] J.S. Eow, M. Ghadiri, A.O. Sharif, Experimental studies of deformation and
break-up of aqueous drops in high electric elds, Colloids and Surfaces A:
Physicochemical and Engineering Aspects 225 (13) (2003) 193210.
[26] J.S. Eow, M. Ghadiri, Dropdrop coalescence in an electric eld: the effects of
applied electric eld and electrode geometry, Colloids and Surfaces A: Physicochemical and Engineering Aspects 219 (2003) 253279.
[27] M. Ghadiri, C.M. Martin, P.A. Arteaga, U. Tuzun, B. Formisani, Evaluation of the
single contact electrical clamping force, Chemical Engineering Science 61 (7)
(2006) 22902300.
[28] J. Sjoblom, Emulsions and Emulsion Stability (Surfactant Science), Marcel
Dekker Inc., New York, 1996.

You might also like